The following article is Open access

Extremely Inclined Orbit of the S-type Planet γ Cep Ab Induced by the Eccentric Kozai–Lidov Mechanism

and

Published 2022 October 7 © 2022. The Author(s). Published by the American Astronomical Society.
, , Citation Xiumin Huang and Jianghui Ji 2022 AJ 164 177 DOI 10.3847/1538-3881/ac8f4c

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

1538-3881/164/5/177

Abstract

γ Cep Ab is a typical S-type planet, which occupies a nearly perpendicular planetary orbit relative to the binary. Here, we use a Markov Chain Monte Carlo sampler to conduct a full N-body fitting and derive self-consistent orbital solutions for this hierarchical system. We then employ the eccentric Kozai–Lidov (EKL) mechanism to explain the extremely inclined orbit of the S-type planet γ Cep Ab. The EKL mechanism plays an essential part in our exploration of the significant oscillations of the mutual inclination imut between the planet and the secondary star. We perform a qualitative analysis and use extensive numerical integrations to investigate the flip conditions and timescales of γ Cep Ab's orbit. When the planetary mass is 15 MJup, the planet can reach imut ∼ 113° given the critical initial conditions of imut < 60° and e1 < 0.7. The timescale for the first orbital flip decreases with the increase of the perturbation Hamiltonian. The flipping orbits of γ Cep Ab are confirmed to have a large possibility of remaining stable, based on surfaces of section and the secular stability criterion. Furthermore, we extend the application of EKL to general S-type planetary systems with a1/a2 ≤ 0.1, where the most intense excitation of imut occurs when a1/a2 = 0.1 and e2 ∼ 0.8, and the variation in planetary mass mainly affects the flip possibility where e1 ≤ 0.3.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

As of today, more than 200 exoplanets have been discovered in binary systems, consisting of circumbinary (P-type) planets and satellite-like (S-type) orbiting bodies (Schwarz et al. 2016). The P-type planets detected by the Kepler Space Telescope are largely coplanar with the binary, where the mutual inclination is less than 2fdg5 (Kostov et al. 2014). The catalog of exoplanets in binary systems 4 reports that only 26% of S-type planets are detected with an orbital inclination of ∼90° through transit observations, while over 50% of this population are observed by radial velocity (RV) without determined inclinations. Therefore, the distribution of the orbital inclinations of planets in binaries plays a significant role in estimating their occurrence rate and understanding their evolution (Armstrong et al. 2014; Gong & Ji 2018).

Figure 1 shows the distribution of the semimajor axes (SMAs) of the planets and secondaries of S-type systems. Here, the blue dashed line denotes an upper limit on the SMA of the secondary, where the perturbation from the secondary is very significant. The red dotted–dashed line indicates the limit on the SMA ratio a1/a2 = 0.1, where a1 and a2 are the SMAs of the planet and secondary, respectively. In particular, three potentially inclined S-type planets, HD 19994 Ab, HD 196885 Ab, and γ Cep Ab, as detected by RV, are labeled in Figure 1.

Figure 1.

Figure 1. Distribution of the SMAs of binaries and planets in S-type systems. The blue dots represent the S-type planets observed by transit, while the red triangles represent those observed by RV. γ Cep Ab is marked by the yellow pentagram.

Standard image High-resolution image

The planet γ Cep Ab is one of the best-known S-type planets in close binary systems. The RV signal of the planet in γ Cep was first measured in 1988 (Campbell et al. 1988), but observation errors and the influence of the secondary star made the planetary detection confusing. Hatzes et al. (2003) revealed the planetary companion to γ Cep A with high-precision RV measurements from 1981 to 2002. The in situ formation of this planet does not seem to be likely, given the truncation model of the protoplanetary disk (Artymowicz & Lubow 1994) and the dynamical instability of the planetesimals in the presence of the star companion γ Cep B (e.g., Jang-Condell et al. 2008; Giuppone et al. 2011). Martí & Beaugé (2012) have suggested that the binary orbital configuration of this system was established after the formation of the planet, as a result of a scattering scenario for the third fly-by star, with γ Cep AB then forming the close binary configuration.

γ Cep Ab occupies a nearly perpendicular orbit relative to the binary (Reffert & Quirrenbach 2011), which presents a tremendous challenge to the migration scenario of a protoplanetary disk with angular momentum exchange. However, close binaries may have a remarkable influence on the formation and evolution of S-type planets, through dynamical perturbations (Xie et al. 2010). Andrade-Ines et al. (2016) studied large parameter spaces of S-type planets to determine the applicability of the disturbing regime up to the second order, including the orbital stability, mean motion resonances, and short-period oscillations. With double stars having a relatively high occurrence rate of 50% (Tokovinin 1997), it would be natural and very likely for a hierarchical triple system hosting an S-type planet to occur, consisting of an inner close binary and a third object in a distant outer orbit.

In the pioneering work of Kozai (1962) and Lidov (1962), secular dynamics were applied to a hierarchical triple system, indicating that an inclined test particle with an inclination of 39° ≤ i ≤ 141° would lead to periodic oscillations of inclination and eccentricity. This is now called the Kozai–Lidov (KL) mechanism. However, the fundamental theories and phenomena of the KL mechanism had been investigated and published by von Zeipel (1910) before the middle of the twentieth century. Afterward, the KL mechanism for hierarchical triple systems was extensively investigated (Harrington 1968; Lee & Peale 2003; Naoz et al. 2013; Teyssandier et al. 2013; Lei et al. 2018; Lei 2019; Tan et al. 2020; Lei 2021) in a wide range of circumstances. Ford et al. (2000) and Eggleton & Kiseleva-Eggleton (2001) showed that the KL mechanism plays an essential part in the secular evolution of hierarchical triple-star systems, and explored the interaction of tidal friction with Kozai cycles in a triple-star system. Such a scenario was then further employed to explain observed close binaries (Fabrycky & Tremaine 2007; Perets & Fabrycky 2009; Thompson 2011; Shappee & Thompson 2013; Naoz & Fabrycky 2014).

Li et al. (2014a) systematically explored the KL mechanism and characterized the parameter space that allows for large-amplitude oscillations in eccentricity and inclination under the test particle limit. As the outer perturber is eccentric and the SMA ratio of the inner orbit and outer orbit meets α = a1/a2 ≪ 1, the polynomial in the classical outer perturbation equation can be expanded to the third order. This allows the octupole-level term, which can give rise to the eccentric KL (EKL) mechanism (Lithwick & Naoz 2011; Naoz 2016). In addition, Li et al. (2014b) confirmed that sufficient initial mutual inclination could produce extremely large eccentricities and flips of the inner orbit, i.e., the orbital inclination transforms between I1 < 90° and I1 > 90° for originally prograde or retrograde orbits.

Additionally, for perturbed orbits with initial low eccentricity and high inclination, the orbital flips induced by the EKL mechanism have been demonstrated to be a kind of resonance, with the libration of a critical angle (Sidorenko 2018). Recently, Lei (2022) systematically studied the EKL mechanism, by analyzing the orbital flipping and its parameterization. The analytical averaging theory interprets the orbit flips as being solutions around polar periodic orbits or kinds of resonant trajectories.

This work aims to explore the inclination evolution and secular stability of potentially inclined S-type planetary systems under the EKL mechanism, particularly for the orbital flip in the γ Cep Ab B system. We conduct a Markov Chain Monte Carlo (MCMC) search using the RV data to derive the best-fitting orbital solution of γ Cep Ab, suggesting that its orbital plane is nearly perpendicular to that of the secondary star. We conclude that such high mutual inclination can provide substantial evidence of orbital flips with different timescales under specific conditions of planetary mass, eccentricity, and mutual inclination, implying that γ Cep Ab may have flipped due to EKL. When the planetary mass is 15 MJup, γ Cep Ab can easily reach the targeted mutual inclination above 120° with critical initial conditions of imut < 60° and e1 < 0.7. Moreover, the flipping cases of γ Cep Ab are proved to be stable from the stability index and Poincaré surfaces of section. Finally, we explore the parameter spaces of planetary mass, SMA, and eccentricities to provide a theoretical prediction when searching for potentially inclined S-type planets in close binary systems. When a1/a2 is fixed, the flip occurs where e1 and e2 are both larger than 0.2, or e1 < 0.2 and e2 > 0.3. The flip likelihood of typical S-type planets is further addressed in order to search for inclined S-type planets induced by EKL (see Figure 1).

This work is structured as follows. In Section 2, the published high-precision RV data are employed to derive the orbital solution of γ Cep Ab, through a full N-body fitting with the MCMC sampler. Section 3 describes the KL and EKL mechanisms. Section 4 provides a qualitative analysis of the numerical results of the maximum inclination and orbital flip timescales under various initial conditions for γ Cep Ab. The stability of the flipping cases is mapped in the planes of (e1, e2) and (e1, g1), as well as the Poincaré surfaces of section. Section 5 extends EKL to more general S-type planets in order to investigate the flip conditions. In Section 6, we summarize the major outcomes.

2. Orbital Solutions of γ Cep Ab

2.1.  N-body Fitting of the RV Data

The γ Cep system is known as a close binary at a distance of 13.79 pc (Hatzes et al. 2003), and it is a candidate target of the CHES mission (Ji et al. 2022). The primary star γ Cep A is a planet-hosting bright star of spectral type K1III–IV, with a stellar mass of m0 = 1.40 ± 0.12 M (Neuhäuser et al. 2007). Neuhäuser et al. (2007) presented a direct detection of the companion γ Cep B, where the parameters of the secondary star are m2 = 0.409 ± 0.018 M, a2 = 20.18 ± 0.66 au, I2 = 119fdg3, Ω2 = 18.04 ± 0fdg98, and the orbital period is P2 = 67.5 ± 1.4 yr. Reffert & Quirrenbach (2011) conducted a fitting only for the inclination I1 and the ascending node Ω1, adopting P, a, e, and K from the best-fitting solution from Butler et al. (2006). They thereby obtained the best-fitting values of I1 = 5fdg7, Ω1 = 37fdg5 or I1 = 173fdg1, Ω1 = 356fdg1 (see Table 1), where I2 and Ω2 are constrained by Neuhäuser et al. (2007).

Table 1. Orbital Solutions of the γ Cep System

ObjectParametersThis WorkHatzes et al. (2003)Torres (2007)Neuhäuser et al. (2007)
   χ2 = 1.48 χ2 = 1.44   
  K1 (m s−1) ${28.08}_{-1.23}^{+1.45}$ ${26.40}_{-1.18}^{+1.41}$ 27.5 ± 1.527.1 ± 1.527.0 ± 1.5
  P1 (days)905.64 ± 2.83 ${901.46}_{-2.83}^{+2.85}$ 905.574 ± 3.08902.8 ± 3.5902.9 ± 3.5
  a1 (AU)2.1459 ± 0.0048 ${2.1376}_{-0.0047}^{+0.0048}$ 2.13 ± 0.051.94 ± 0.062.044 ± 0.057
  e1 ${0.0724}_{-0.0575}^{+0.0879}$ ${0.0856}_{-0.0624}^{+0.0866}$ 0.12 ± 0.050.113 ± 0.0580.115 ± 0.058
γ Cep Ab ω1 (deg) ${48.47}_{-1.81}^{+4.56}$ ${55.37}_{-4.57}^{+8.81}$ 49.6 ± 25.663.0 ± 27.063.0 ± 27.0
  T0,1 (HJD-2400000) ${53140.16}_{-38.31}^{+34.15}$ ${53107.63}_{-30.90}^{+25.47}$ 53121.9 ± 66.953146.0 ± 72.053146.0 ± 71.0
  ${m}_{1}\sin {I}_{1}\left({M}_{\mathrm{Jup}}\right)$ ${1.7420}_{-0.0743}^{+0.0733}$ ${1.6306}_{-0.0694}^{+0.0698}$ 1.70 ± 0.401.43 ± 0.131.60 ± 0.13
 Ω1 (deg)37.5 a 356.1 a .........
  I1 (deg)5.7 a 173.1 a ...>4.9...
  K2 (m s−1) ${1821.70}_{-1.01}^{+0.95}$ ${1699.94}_{-3.19}^{+3.45}$ 1820.0 ± 49.01925.0 ± 14.01932.0 ± 14.0
  P2 (days) ${20754.66}_{-57.09}^{+59.01}$ ${20731.68}_{-59.53}^{+57.18}$ 20750.658 ± 1568.624392 ± 52224654.375 ± 511
  a2 (AU) ${18.6421}_{-0.0381}^{+0.0392}$ ${18.6217}_{-0.0389}^{+0.0381}$ 18.5 ± 1.119.02 ± 0.6420.18 ± 0.66
  e2 0.3603 ± 0.00120.3605 ± 0.00260.3610 ± 0.02300.4085 ± 0.00650.4112 ± 0.0063
γ Cep AB ω2 (deg)158.92 ± 0.20158.90 ± 0.20158.76 ± 1.20160.96 ± 0.40161.01 ± 0.40
  T0,2 (HJD-2400000) ${48435.22}_{-0.37}^{+0.80}$ ${48435.04}_{-0.88}^{+0.46}$ 48429.03 ± 27.048479.0 ± 12.048444.8 ± 11.16
  ${m}_{2}\left({M}_{\odot }\right)$ 0.3991 ± 0.00050.3986 ± 0.0005...0.362 ± 0.0220.409 ± 0.018
 Ω2 (deg)18.04 ± 0.98 b 18.04 ± 0.98 b ...13.0 ± 2.418.04 ± 0.98
  I2 (deg)119.3 ± 1.0 b 119.3 ± 1.0 b ...118.1 ± 1.2119.3 ± 1.0

Notes.

a Astrometric data fitting results adopted from Reffert & Quirrenbach (2011). b Orbital solutions adopted from the imaging observations in Neuhäuser et al. (2007).

Download table as:  ASCIITypeset image

By using the most recent spectroscopic observations from the Canada–France–Hawaii Telescope (CFHT; Walker et al. 1992) and the McDonald Observatory Planetary Search (MOPS) program (Hatzes et al. 2003), two-Keplerian solutions of the γ Cep system have previously been given (Hatzes et al. 2003; Neuhäuser et al. 2007; Torres 2007). In these studies, an uncoupled two-Keplerian orbital fitting approach is utilized with the standard nonlinear least-squares technique. However, for a hierarchical system with a massive secondary companion, the significant mutual interaction between the planet and the star in the pair is supposed to be considered under the Jacobi frame, in order for the real dynamics to be represented (Lee & Peale 2003).

Here, we solve the orbits of the planet and the star companion with the MCMC ensemble sampler emcee (Foreman-Mackey et al. 2013) in the full N-body model (Ford 2006; Nelson et al. 2016). 14 parameters of $\{{m}_{1}\sin {I}_{1}$, P1, $\sqrt{{e}_{1}}\sin {\omega }_{1}$, $\sqrt{{e}_{1}}\cos {\omega }_{1}$, M1, ${m}_{2}\sin {I}_{2}$, P2, $\sqrt{{e}_{2}}\sin {\omega }_{2}$, $\sqrt{{e}_{2}}\cos {\omega }_{2}$, M2}, plus the RV offsets {RV0,1, RV0,2, RV0,3, RV0,4} of four time series (CFHT, MOPS I, MOPS II, and MOPS III), are adopted for fitting at the first observation epoch (HJD-2444754.129) in the Jacobi reference frame (Lee & Peale 2003). $\sqrt{e}\sin \omega $ and $\sqrt{e}\cos \omega $ are more efficient than e and ω for low-eccentricity planets, and can avoid a situation of multiple solutions (Ford 2006).

We then utilize the N-body integrator IAS15 (Rein & Spiegel 2015), which is an integrator with an adaptive step-size control, to compute the perturbed orbits of the planet and the secondary star in each step of the emcee sampler. The initial orbital inclinations and ascending nodes are assumed to be known (Reffert & Quirrenbach 2011): I1 = 5fdg7, Ω1 = 37fdg5 or I1 = 173fdg1, Ω1 = 356fdg1. Other initials include the planetary mass m1 and the secondary mass m2, the SMA a1,2, the eccentricity e1,2, and the argument of periastron ω1,2, which can be derived from the resultant fitting parameters. The integration precision is given as 10−15 and the minimum time step is set to 0.001 days. The theoretical RV signals of the observation epochs are calculated from the Jacobian orbital elements, with the RV semi-amplitude K1,2 being defined in Lee & Peale (2003):

Equation (1)

where P1,2, I1,2, and e1,2 are time-variable orbital elements at each observation epoch.

To derive the N-body best-fitting solutions for this system, we run 100,000 steps each for 28 walkers in the 14-dimensional parameter space. The mean acceptance fraction of the sampler walkers is 0.36, which is in the range from 0.2 to 0.5 that is suggested to produce representative samples (Foreman-Mackey et al. 2013), and the autocorrelation time for each fitting parameter ranges from 200 to 500 steps, which is smaller than our number of sampling steps to ensure the convergence of the MCMC chains. Here, the autocorrelation time represents the number of steps required for the chain to produce an independent sample (Foreman-Mackey et al. 2013).

Finally, we report two set of solutions of the minimum mass $m\sin I$, the RV semi-amplitude K, the SMA a, the eccentricity e, the argument of periastron ω, and the epoch of periastron passage T0 in Table 1, with χ2 = 1.48 for I1 = 5fdg7 and χ2 = 1.44 for I1 = 173fdg1, respectively. In Figures 2 and 3, we plot the best-fitting RV signals from the N-body solutions (the black solid lines), using the measurements versus the epoch with χ2 = 1.44, with respect to ${{RV}}_{\mathrm{0,1}}={1290.0}_{-17.6}^{+18.5}$ ms−1, ${{RV}}_{\mathrm{0,2}}={2035.2}_{-18.8}^{+20.3}$ ms−1, ${{RV}}_{\mathrm{0,3}}={2231.1}_{-16.8}^{+19.2}$ ms−1, and ${{RV}}_{\mathrm{0,4}}={865.0}_{-19.4}^{+22.4}$ ms−1, respectively.

Figure 2.

Figure 2. The best-fitting RV signals of the γ Cep Ab B system. Upper panel: the red triangles and squares show published observations (Walker et al. 1992; Hatzes et al. 2003), where the black solid line denotes the RV signals of the N-body model, with mutual perturbation between the companion and planet b. Lower panel: O-C for the N-body model.

Standard image High-resolution image
Figure 3.

Figure 3. The blue dots with red error bars are the residuals after subtracting the calculated RV of γ Cep AB from the observations (O-AB). The derived RV signals induced by γ Cep Ab from the N-body fitting are shown by the black solid curve, where all the orbital elements change over time.

Standard image High-resolution image

By subtracting the theoretical RV induced by the secondary, Figure 4 presents the Lomb–Scargle periodogram analysis of the planet b, where the strongest signal reveals an orbital period of roughly 900 days. Figure 5 shows the posterior distributions of $\{{m}_{1}\sin {I}_{1}$, P1, $\sqrt{{e}_{1}}\sin {\omega }_{1}$, $\sqrt{{e}_{1}}\cos {\omega }_{1}$, M1, ${m}_{2}\sin {I}_{2}$, P2, $\sqrt{{e}_{2}}\sin {\omega }_{2}$, $\sqrt{{e}_{2}}\cos {\omega }_{2}$, M2}. We report the median (50th percentile) of the posterior distribution as the best-fitting value, with the 1σ uncertainties being provided using the 16th and 84th percentiles.

Figure 4.

Figure 4. Lomb–Scargle periodogram analysis of γ Cep Ab. The value of the power is the maximum power of the periodic signals, while the dashed lines of false alarm probability (FAP)1, FAP2, and FAP3 represent 1%, 0.1%, and 0.01% of the FAP, respectively. There is a significant periodic signal of roughly 900 days, which means that this signal is truly periodic.

Standard image High-resolution image
Figure 5.

Figure 5. Corner diagram of our N-body best-fitting solutions, with χ2 = 1.44. Here, $m\sin I$, P, $\sqrt{e}sin{\omega }$,$\sqrt{e}\cos \omega $, and M with subscripts 1 and 2, denote the minimum mass, orbital period, $\sqrt{e}$ vector, and mean anomaly of γ Cep Ab and γ Cep AB, respectively. The figure illustrates the one- and two-dimensional projections of the posterior probability distributions of the parameters, with the structure directly showing the covariances between the two parameters. The distribution histograms of each parameter are listed on the diagonal.

Standard image High-resolution image

2.2. Estimation of the Planetary Mass

To derive more reliable orbital solutions, it is necessary to utilize high-precision RV and astrometric measurements in the fitting procedure at the same time. As is well known, Gaia Data Release 2 (Gaia Collaboration et al. 2018) was first accessible in 2018, and the science team recently announced Gaia Early Data Release 3 (Gaia Collaboration et al. 2021). For most of the sources in the Gaia catalogs, sequential astrometric data are not yet available, and will be released in Gaia Data Release 4. Thus, improvements of the planetary masses from Gaia will be expected in the future.

Table 1 summarizes our derived orbital parameters of the γ Cep system. The minimum planetary mass is fitted as ${1.7420}_{-0.0743}^{+0.0733}$ MJup or ${1.6306}_{-0.0694}^{+0.0698}$ MJup. Note that the real mass of planet b is significantly dependent upon the accuracy of the RV observations and the dynamical integrations; here, we simply estimate the mass range of γ Cep Ab. When the observed inclination of the planet is I1 ∈ [3fdg8, 20fdg8] (Reffert & Quirrenbach 2011), we obtain an estimated planetary mass m1 ∈ [5.0, 26.6] MJup, while for I1 ∈ [166fdg6, 174fdg8] (Reffert & Quirrenbach 2011), m1 ∈ [7.1, 26.2] MJup can be derived. In addition, based on the planetary mass limit of 16.9 MJup (Torres 2007), we here assume m1 ∈ [5, 16.9] MJup.

To determine the current observed mutual inclination imut in the γ Cep system, we use the law of cosines for angles of a spherical triangle (Gellert et al. 1977):

Equation (2)

deriving imut = 113fdg9 for I1 = 5fdg7. Thus, a question naturally arises—what kind of dynamical process could trigger such high mutual inclinations of the two bodies, and could the EKL mechanism play a role in the secular evolution of γ Cep Ab? In the following section, we describe the EKL mechanism and explore the secular evolution of γ Cep Ab under EKL.

3. Dynamical Model

3.1. KL Mechanism

In the secular evolution of triple-body systems, the perturbation arising from the third body acts on a timescale much longer than its orbital period. One typical origin of secular resonance is from the perturbation potential from adjacent orbits. When the mass of the perturbed body is negligibly small in comparison to those bodies in a hierarchical system, the test particle approximation comes into play. At this time, if the perturbation of a circular outer orbit works, the Hamiltonian of this system can be expanded to the quadrupole level, known as the KL mechanism, as aforementioned.

The KL mechanism addresses the inclination and eccentricity of the inner test particle oscillating over secular evolution in a hierarchical system, where the test body and the primary star are surrounded by a distant companion (von Zeipel 1910; Kozai 1962; Lidov 1962). With the secular approximation, the inner and outer orbits only exchange angular momentum, so the SMAs of the orbits do not change. When the SMA ratio αr = a1/a2 is a small parameter, the perturbation term of the complete Hamiltonian can be expanded as a power series in αr (Naoz 2016):

Equation (3)

Equation (4)

where m0 is the mass of the primary, m1 and m2 are the masses of the inner and the outer body, k2 is the gravitational constant (with the mass unit of M and the length unit of au), r1 is the distance between m0 and m1, r2 is the distance between the center of mass of the inner binary and m2, Pj is the Legendre polynomial, and Φ is the angle between the vectors r1 and r2 (the subscripts j = 1, 2 represent the inner and outer orbits, respectively).

The invariable plane reference frame is imported here to define three important Delaunay elements (Valtonen & Karttunen 2006): l, g, h, and their conjugate momenta L, G, H, where l, g, and h are the notations of the mean anomaly M, the argument of periastron ω, and the longitude of ascending node Ω, respectively. As shown in Figure 6, G tot is the total angular momentum vector of the system, which is conserved over the secular evolution.

Figure 6.

Figure 6. Geometry of the invariable plane and the reference plane system. G1 and G2 are angular momentum vectors of the inner and outer orbits; Gtot is the total angular momentum vector of the system, with the invariable plane being perpendicular to Gtot; i1,2 is the angle between the orbital angular momentum vector and G tot, where imut is the mutual inclination between the inner and outer orbit; and I1,2 represents the observed inclination of the inner and outer orbit, which is the angle between the orbital plane and reference plane (the sight plane).

Standard image High-resolution image

The three conjugate momenta L, G, H are expressed as (Naoz et al. 2013):

Equation (5)

where L is only determined by constant parameters, including the masses m0, m1, and m2, the SMAs a1, a2, and the gravitational constant k2. Thus, L is a constant for a specific system in the evolution, while G and H are time-varying. G represents the magnitude of each orbit's angular momentum and H is the component of G along the z-axis.

According to the geometric relations and the assumption of h1h2 = π, the mutual inclination between the inner and outer orbit imut could be derived as (Naoz et al. 2013):

Equation (6)

Generally, the equations of motion can be expressed by canonical relations of the three conjugate momenta, L, G, H, and the three Delaunay elements, l, g, h, as the mean anomaly can be eliminated under the double-averaged secular approximation, and h1 and h2 in equations of motion have been removed by the relation h1h2 = π. The time evolution for ω, e, and i can easily be derived from the reduced canonical relations (Naoz 2016):

Equation (7)

where j = 1, 2. The original nonplanar three-body system can thus be reduced to a dynamical system of 2 degrees of freedom (dof).

In Naoz et al. (2013), i2 is set to be 0, thus imut is equal to i1. In this work, we treat the orbital plane of the secondary as the invariable plane, where imut = i1. In the following sections, we redefine the orbital flip of the planet b as the variation of imut around 90°, instead of the real observed orbital inclination.

3.2. EKL Mechanism in Nonrestricted Triple Systems

In the γ Cep Ab B system, e2 is close to 0.36 and the estimated maximum mass of the planet is near the deuterium-burning limit. Thus, γ Cep Ab B is a nonrestricted triple system with an eccentric outer orbit, where the octupole-level terms in the Hamiltonian can become important, and the eccentricities of the two orbits are coupled and oscillate simultaneously over secular evolution. Naoz et al. (2013) called it the EKL mechanism in nonrestricted triple systems and derived the complete Hamiltonian of the system, including the octupole term in addition to the quadrupole term:

Equation (8)

where

Equation (9)

Equation (10)

Equation (11)

Equation (12)

Equation (13)

The secular perturbation theory in this specific case is called the EKL mechanism, as previously mentioned. The time evolution for ω, e, and imut can be derived through Equation (8) as well. Here, epsilonM can be used as an indicator to characterize the strength of the octupole-level effect on the system's perturbation potential.

4. Secular Evolution of γ Cep Ab

As we described in Section 1, the EKL mechanism plays a crucial role in the secular evolution of celestial bodies. Here, in the γ Cep Ab B system, the inner planet b is assumed to be a substellar object with a remarkable mass (Reffert & Quirrenbach 2011); thus, the test particle approximation is not applicable. With a1 = 2.1376 au, a2 = 18.6217 au, and the SMA ratio of planet b and the secondary αr ∼ 0.1, the perturbation term of the complete Hamiltonian can thus be expanded as a power series in αr . As the eccentricity of the outer orbit e2 ∼ 0.4, the EKL mechanism can play a significant part when exploring the secular evolution of the γ Cep Ab B system. Here, we will investigate the mutual inclination oscillations of the γ Cep Ab B system to examine whether planet b can achieve an extremely inclined orbit through EKL.

In this section, we are mainly concerned with the flip conditions and timescale for the first flip of i1, which is equal to imut. The stability of the rolling-over orbits is further explored. To investigate the relative global dynamics of the system, we show several kinds of representative planes of initial conditions. In Section 4.1, the (e1, imut) plane is first used to search the flip conditions. In Section 4.3, in order to obtain the global qualitative structure of the averaged 2 dof Hamiltonian system, we import the representative plane of (e1, e2) (Michtchenko & Malhotra 2004) and the phase space of (e1, g1) (Tan et al. 2020). We then apply the ($\sqrt{1-{e}_{1}^{2}}$, $\sqrt{1-{e}_{2}^{2}}$) plane as the representative plane for studying the relative global dynamical features of the γ Cep system, and employ the (e1, g1) plane to theoretically define quasiperiodic and circulating orbits.

4.1. Orbital Flip Conditions

The investigation of the amplitude of the inclination oscillation reveals that γ Cep Ab could evolve into an extremely inclined orbit. However, one may encounter difficulties in accurately predicting the theoretical correlation between the extent of the orbital inclination excitation and the initial conditions. Katz et al. (2011) defined a complex function of the analytical critical flip condition by averaging over the quadrupole-level effect.

We provide the planetary mass m1 ∈ [5, 16.9] MJup in Section 2.2. Here, we assume the mass of γ Cep Ab to be m1 = {5, 9, 11, 15} MJup, then calculate the ranges of the perturbation Hamiltonian ${ \mathcal H }$ and the total angular momentum Gtot for each value of m1 with known parameters {m0, m2, a1, a2} and e1 ∈ [0, 1], e2 ∈ [0.35, 0.45], and imut ∈ [0°, 180°]. Furthermore, we set g1 = 0° and g2 = 0° in the initial conditions, since g1 and g2 can always go through either 0° or 180° over secular evolution.

The contour maps of the total angular momentum Gtot and the perturbation Hamiltonian ${ \mathcal H }$ are simultaneously plotted in the (e1, imut) plane. In Figure 7, we show two examples of m1 = {5, 15} MJup with e2 ∈ [0.35, 0.45], which is around the current observation and will not vary significantly over the evolution. For other planetary masses, the structures of the ${ \mathcal H }$Gtot contour maps have similar features with different ranges of Gtot and ${ \mathcal H }$. When m1 = 5 MJup, Gtot ∈ [1.7050, 1.8025], ${ \mathcal H }\in [-1.05\times {10}^{-6}$, 3.0 × 10−7], and when m1 = 9 MJup, Gtot ∈ [1.6994, 1.8137], ${ \mathcal H }\in [-1.75\times {10}^{-6}$, 5.0 × 10−7]. When m1 = 11 MJup, Gtot ∈ [1.6975, 1.8187], ${ \mathcal H }\in [-2.10\times {10}^{-6}$, 6.0 × 10−7], and when m1 = 15 MJup, Gtot ∈ [1.696, 1.824] and ${ \mathcal H }\in [-2.8\times {10}^{-6}$, 8.0 × 10−7]. The cross points of the Gtot and ${ \mathcal H }$ contours indicate all the possible initial conditions of imut and e1 over the secular evolution. The range of values of Gtot and ${ \mathcal H }$ in Figure 7 will change with variational planetary masses, while the structure of the contours will remain similar.

Figure 7.

Figure 7. The contour maps of the total angular momentum Gtot and the perturbation Hamiltonian ${ \mathcal H }$ for m1 = 5 and 15 MJup in the initial conditions plane of (e1, imut). The white dashed lines represent the constant values of Gtot for e2 ∈ [0.35, 0.45], while the color bars on the right indicate the different values of the perturbation Hamiltonian. In panel (a), when e2 selects five different values, there will be five level curves of Gtot. We also mark out the initial conditions for Gtot = 1.75318 (the black dashed lines) and ${ \mathcal H }=-8.06\times {10}^{-8}$ (the solid line) with black dots, corresponding to data points with red error bars in Figure 8(a).

Standard image High-resolution image

To further derive the evolution results of these general initial conditions, we uniformly choose specific values of ${ \mathcal H }$ and Gtot between the upper and lower limits in Figure 7. For a given ${ \mathcal H }$, based on the conserved Hamiltonian and the total angular momentum shown in Equations (5) and (8) of the octupole perturbation theory, we choose at least 40 cases of initial e1 and imut. Considering the conservation of energy and the total angular momentum, we further explore the extreme value of the orbital inclination and orbital stability under these selected initial conditions in detail.

In the non–test particle approximation under the classical KL mechanism, the eccentricity and inclination oscillate regularly over a well-defined timescale tquad (Antognini 2015):

Equation (14)

This relationship was derived under consideration of the equation of motion of ω, by integrating between the maximum and minimum eccentricities. Here, tquad can be applied to estimate the timescale in the EKL scenario.

According to the values of m0, m1, m2, e2, a1, and a2, the quadrupole period tquad of the γ Cep system is estimated to be ∼1000 yr, which is consistent with our numerical simulation results. Here, we investigate the secular evolution of γ Cep Ab by considering a diverse planetary mass and performing the simulation for 100 Myr (∼105 tquad) using the RKF7(8) integrator. The observed orbital inclinations I1 and I2 are required to calculate the constant total angular momentum and Hamiltonian. Hereafter, i1 and i2 denote the angles of the orbital plane relative to the invariable plane. However, the true orbital inclinations of the inner and outer orbits over the evolution are not well known.

Panels (a)–(d) of Figure 8 each plots three sets of typical outcomes, with ${{ \mathcal H }}_{1}$, ${{ \mathcal H }}_{2}$, and ${{ \mathcal H }}_{3}$. If m1 = 5 MJup, both the prograde and retrograde orbits will flip when e1 < 0.5 for ${{ \mathcal H }}_{1}=-8.06\times {10}^{-8}$, e1 > 0.65 for ${{ \mathcal H }}_{2}=-4.95\times {10}^{-7}$, and e1 > 0.8 for ${{ \mathcal H }}_{3}=-6.08\times {10}^{-7}$. We note that most intense orbital flips take place when the initial imut is close to 90°.

Figure 8.

Figure 8. Ranges of imut for m1 = 5, 9, 11, and 15 MJup over a timescale of 100 Myr. The black dots represent the positions of the initial parameters {e1,0, imut,0}. The upper and lower limits of the error bars represent the maximum and minimum values of the mutual orbital inclination imut, respectively. The different colors correspond to specific values of the perturbation Hamiltonian.

Standard image High-resolution image

From the estimated ${m}_{1}\sin {I}_{1}$ in Table 1, and the law of cosines for the angles of a spherical triangle in Equation (6), if m1 = 5 MJup and the observed inclination of planet b is roughly 20°, the mutual inclination imut ∼ 100° is I2 = 119fdg3. As a consequence, it is clear that under the selected values of ${{ \mathcal H }}_{1}$, ${{ \mathcal H }}_{2}$, and ${{ \mathcal H }}_{3}$, such a high mutual inclination could always be achieved for original prograde orbits. In fact, as shown in Figure 8, the evolution of the mutual inclination of prograde and retrograde orbits is approximately symmetric about 90°, which can also be obtained from Equation (8).

Similarly, if m1 = 9 MJup, the observed inclination of planet b is nearly 11°, thus the mutual inclination imut is estimated to be 108°. When ${{ \mathcal H }}_{1}=-1.45\times {10}^{-7}$, e1 ∼ 0.5, ${{ \mathcal H }}_{2}=-6.35\times {10}^{-7}$, e1 ∼ 0.65, and ${{ \mathcal H }}_{3}=-1.22\times {10}^{-6}$, e1 ∼ 0.8, the amplitude of inclination for the orbital flip could be raised up to about 180°.

When m1 is equal to 11 MJup, the calculated inclination of γ Cep Ab is about 9° and the targeted mutual inclination imut ∼ 110° can be achieved when ${{ \mathcal H }}_{1}=-1.98\times {10}^{-7}$, with e1 ∼ 0.5, ${{ \mathcal H }}_{2}=-8.56\times {10}^{-7}$, with e1 ∼ 0.7, and ${{ \mathcal H }}_{3}\,=-1.7\times {10}^{-6}$, with e1 ∼ 0.85. Thus, the critical value of the initial eccentricity for the orbital flips of ${{ \mathcal H }}_{2}$ in Figure 8(c) is larger than the values for Figures 8(a), (b), and (d).

When m1 is 15 MJup, the observed inclination of γ Cep Ab is about 6° and the critical conditions for the targeted mutual inclination imut ∼ 113° are ${{ \mathcal H }}_{1}=-2.4\times {10}^{-7}$, with e1 ∼ 0.5, ${{ \mathcal H }}_{2}=-1.44\times {10}^{-6}$, with e1 ∼ 0.65, and ${{ \mathcal H }}_{3}=-2.03\times {10}^{-6}$, with e1 ∼ 0.8. In Figure 8(d), for ${{ \mathcal H }}_{2}=-1.44\times {10}^{-6}$, the orbital flip occurs when the initial imut < 60°. Thus, it is easier for γ Cep Ab to reach the targeted mutual inclination with m1 = 15 MJup, e1 < 0.7, and the critical initial imut being lower than 60°.

Above all, we conclude that the initial conditions for the orbital flips under investigation for the γ Cep system are e1 > 0.5 and imut ∈ [60°, 120°], and that various planetary masses simply affect the critical eccentricity when e1 > 0.6, with little influence on the maximum mutual inclination. The distribution tendency of the flipping conditions with various initial inclinations and eccentricities is similar to that in Lei (2022), while we perform the variation of the flipping conditions under different m1, which affects the octupole-level factor epsilonM .

4.2. Orbital Flip Timescale

To obtain an in-depth understanding of the rolling-over orbits reported in Section 4.1, we will further explore the orbital flip timescale over secular evolution, which may rely on the initial conditions, e.g., m1, e1, and imut. The duration of the flip is critical for the observational possibility of potentially extremely inclined S-type planets that are transforming between prograde and retrograde orbits. An approximate analytical timescale for the first flip for a nonchaotic orbit when m1 → 0 has been given by Katz et al. (2011) and Antognini (2015):

Equation (15)

where

Equation (16)

This theoretical flip timescale is effective when the initial conditions meet e1 → 0, ω1 → 0, and imut → 90°. Note that the flip timescale for the circular test particle approximation is mainly determined by e2 when the object masses and the orbital SMAs are fixed. Nevertheless, in the case of high-inclination oscillation, the timescale for the first flip is difficult to quantitatively assess, because this evolution is likely to be chaotic. In order to investigate the dependence of the orbital flip possibility on the integration timescale, we adopt the definition of the flip ratio f = tflip/ttotal (Teyssandier et al. 2013) in order to describe the timescale for the first flip and the observational possibility of the orbital flip process over secular evolution, where tflip is the duration of the orbital flip process and ttotal represents the total evolution timescale. We extract the cases from Section 4.1 in which orbital flips occur in Figure 9. For these rolling-over orbits, the integrations stop when e1 becomes larger than 0.9999 and the planet falls into the Roche limit of the primary star.

Figure 9.

Figure 9. The evolution of the flip ratio for m1 = 5, 9, 11, and 15 MJup over a timescale of 100 Myr. As in Figure 8, ${{ \mathcal H }}_{1}$, ${{ \mathcal H }}_{2}$, and ${{ \mathcal H }}_{3}$ correspond to the selected perturbation Hamiltonian. In each panel, the evolution before 2 Myr is given as a subfigure. In these subfigures, dotted–dashed lines with an equilibrium value f > 0.5 represent prograde orbits that have been transformed from retrograde orbits, while solid lines with f < 0.5 have been transformed from prograde orbits to retrograde orbits.

Standard image High-resolution image

In Figure 9(a), we observe that when ${{ \mathcal H }}_{1}=-8.06\times {10}^{-8}$, the orbits turn over after 30 Myr and take more than 80 Myr to arrive at the equilibrium f. However, the flips under ${{ \mathcal H }}_{2}$ and ${{ \mathcal H }}_{3}$ occur from the very beginning of the evolution, and reach the equilibrium value before 1 Myr. This phenomenon also shows up in Figures 9(b) and (c), indicating that the flip timescale decreases with the increase of ${ \mathcal H }$, under most circumstances. The oscillation timescale of f over the flip procedure decreases with the rise of the perturbation Hamiltonian, as can be seen from the blue and green curves in the four subfigures.

In Figure 9(d), there is a special flipping case for ${{ \mathcal H }}_{2}$ when m1 = 15 MJup. This rolling-over orbit has initial conditions of e1 = 0.64 and imut = 58fdg9 in Figure 8(d). In comparison to the other cases under ${{ \mathcal H }}_{2}$ for different masses of planet, it can be concluded that a relatively low initial imut below 60° leads to a relatively large timescale for the first orbital flip, as well as more time to reach the equilibrium value of the flip ratio.

Moreover, we investigate the equilibrium value f for the flip cases for ${{ \mathcal H }}_{2}$ and ${{ \mathcal H }}_{3}$. We find that the final f of the dotted–dashed lines are all located above 0.5, while those of the solid lines are all lower than 0.5. Thus, the time duration of the flip process for those original retrograde orbits is larger. The maximum f occurs in Figure 9(b), with f ∼ 0.65 when m1 = 9 MJup, ${{ \mathcal H }}_{2}=-6.35\times {10}^{-7}$, whereas the minimum f occurs in Figure 9(d), with f ∼ 0.4 when m1 = 15 MJup, ${{ \mathcal H }}_{3}=-2.03\times {10}^{-6}$. Thus, the equilibrium value of the flip ratio under the EKL mechanism is related to the planetary mass, initial eccentricities, and mutual inclinations. When the equilibrium value of f is closer to 0.5, the observational possibility of the retrograde orbit transforming from the prograde is higher.

4.3. Stability of Flipping Orbits

For a planetary system with given orbital parameters, the periodic orbits and stability are identified by the representative plane of (e1, e2), the level curves in the (e1, g1) plane, the Poincaré surface of section, and the long-term stability criterion. These methods are applied to our investigation of the stability of specific S-type planets in binary systems, while the perturbative treatment and the invariant manifolds characterize the stability with a fixed Hamiltonian (Lei 2022).

To derive a global and comprehensive view of the system dynamics, we first attempt to construct a parametrical analysis of the secular model. We first show the representative plane of (e1, e2) by Michtchenko & Malhotra (2004), which was then followed by studies in secular dynamics and resonances (Libert & Henrard 2006, 2007; Henrard & Libert 2008; Libert & Henrard 2008). This approach relies on the secular averaging of the full Hamiltonian and can be applied to construct global phase portraits of the systematic variables, thereby detecting resonances and discerning periodic orbits in the N-body dynamics.

According to Michtchenko & Malhotra (2004), the secular motion of planetary systems is mainly decided by the global quantities of total energy and the angular momentum deficit. The phase structures are determined by two constants: a1/a2 and m1/m2. The description of the secular behavior of this system is derived by the distribution of the initial values of e1, e2, g1, g2, and imut. To simplify the model, we fix m1 = 15 MJup and imut = 60° in Figure 10, $\sqrt{1-{e}_{1}^{2}}$ cosg1 > 0 with g1 = 0°, while $\sqrt{1-{e}_{1}^{2}}$ cosg1 < 0 with g1 = 180°, and g1 can always go through 0° and 180° over the evolution. As seen from Figure 10, we may come to the conclusion that periodic orbits occur when Gtot and ${ \mathcal H }$ have two cross points.

Figure 10.

Figure 10. One of the contour maps of the total angular momentum and the perturbation Hamiltonian in the representative plane of ($\sqrt{1-{e}_{1}^{2}}$, $\sqrt{1-{e}_{2}^{2}}$). The black dashed lines indicate the level curves of the Hamiltonian and the red solid line is the conserved Gtot.

Standard image High-resolution image

For the secular evolution of the γ Cep Ab B system, one couple of free variables is e1 and g1 in the 2 dof averaged problem. Thus, we can first easily perform the qualitative analysis in the (e1, g1) plane (Tan et al. 2020). Given the limit of Gtot in Section 4.1, we select and fix the value of Gtot to present level curves of the Hamiltonian in the parameter space of (e1, g1), as shown in Figure 11. The green stream lines for circulating orbits and the blue circles for resonant orbits are easy to distinguish. With the changes of the perturbation Hamiltonian, the dynamical structure transitions from circulation to libration and the evolution interval of e1 vary. The smallest circles have the largest magnitude of ${ \mathcal H }$, which indicates that the secular resonance effect is the most intense, as shown in Figure 12, with further details.

Figure 11.

Figure 11. Two examples of Hamiltonian level curves with m1 = 11 MJup and 15 MJup in the (e1, g1) plane, where e2 and Gtot are fixed. The blue circles and green curves are resonant and oscillating orbits with equilibrium points of 90° and 180°, respectively.

Standard image High-resolution image
Figure 12.

Figure 12. Surfaces of sections in the g1 = 0 plane, with different masses of γ Cep Ab, by varying the perturbation Hamiltonian ${ \mathcal H }({\rm{\Delta }}h\to \pi )$. Panels (a), (c), and (e): dynamical maps for m1 = 11 MJup. Panels (b), (d), and (f): dynamical maps for m1 = 15 MJup.

Standard image High-resolution image

In addition to the representative plane of ($\sqrt{1-{e}_{1}^{2}}$, $\sqrt{1-{e}_{2}^{2}}$) and the level curves in the (e1, g1) plane, we further investigate the stability of the rolling-over orbits under the selected Hamiltonian with the Poincaré surface of section. For a two-dimensional Hamiltonian system, the Poincaré surface of section is a commonly used method: under a given energy integral, the phase flow of the system is three-dimensional. By selecting a suitable section, the intersection point of the systematic phase flow and the section is projected in two dimensions. In other words, the Poincaré surface of section reflects the geometry of the system dynamics. Given the integration of energy (i.e., the octupole perturbation Hamiltonian) and the selection of the section (i.e., the g1 = 0 section), the surface of section can be generated by plotting intersection points in the g2e2 frame for the plane of g1 = 0.

The surface of section enables us to identify the order of orbital resonance and the dynamical stability of the system through a series of geometric structures. For a given fixed value of the Hamiltonian, the orbital mode can be derived from only two orbital parameters. Various orbital modes, including resonance, circulation, and chaos, are distinguished by the shape of the point distribution in the section.

We present typical structures in the surfaces of section of the γ Cep Ab B system in Figure 12. The surfaces of section are plotted in the g1 = 0 plane for m1 = {5, 9, 11, 15} MJup, with a corresponding perturbation Hamiltonian. To initially set e2 and g2 in the uniform grid, we solve other initial parameters, based on the conservation of the system's energy and the total angular momentum. There are two kinds of regions in these plots: the closed circles are resonant orbits (quasiperiodic orbits), where e2 and g2 oscillate in the bounded regions, while the stream lines denote circulative orbits, where at least one of the parameters circulates.

Figure 12 simply exhibits the dynamical structure of the surfaces of section for m1 = 11 and 15 MJup. With increasing ${ \mathcal H }$, the structures in the surfaces of section appear to be diverse. As shown in Figures 12(a), (c), and (e), when m1 = 11 MJup, the orbits with respect to g2 = 180° disappear, whereas both orbits with fixed points of g2 = 0° and 180° remain. These libration regions perfectly match the orbits that would turn over. In Figures 12(b), (d), and (f), when m1 = 15 MJup, the positions of the fixed angles of g2 are the same as for m1 = 11 MJup. Figure 12(a) shows that the orbital flips of γ Cep Ab are dominated by the octupole-level resonance, and the oscillation amplitude of e2 in these flipping cases is close to 0.01. In Figures 12(e) and (f), when m1 = 11 MJup, ${ \mathcal H }=-1.7\times {10}^{-6}$, and m1 = 15 MJup, ${ \mathcal H }=-2.03\times {10}^{-6}$, respectively, most of the orbits around the initial values of e2 could turn over in the secular evolution. Additional simulations for m1 = 5 and 9 MJup reveal similar structures, but with various oscillation intervals of e2, when compared to above results.

The orbital flip criterion of e2 can be evaluated with the conserved total angular momentum in Equation (5), when i2 = 0° and imut = 90°:

Equation (17)

We mark this flip criterion in Figure 12 with the horizontal red lines. The circles and stream lines crossing the horizontal red line represent those orbits that invert regularly, while the ellipses above the horizontal red line in Figures 12(a)–(b) stand for the orbits that do not invert. Here, we find that the structures in the phase portrait become much clearer with the decrease in the perturbation Hamiltonian ${ \mathcal H }$, where the chaotic orbits disappear, similar to those in Lei (2022).

Additionally, we compare the boundaries of the regions of orbital flips in the (e1,0, imut,0) space for the nonrestricted model with those of Figure 3 (Lei 2022). By analyzing the results in the Poincaré surfaces of section, the inner orbit appears to be more stable in our hierarchical system when the planetary eccentricity 0.5 < e1,0 < 0.6, which is simply occupied by circulating and librating orbits.

In order to further confirm whether the orbital flip in γ Cep Ab is stable, we adopt the long-term stability criterion given by Mardling & Aarseth (2001):

Equation (18)

where e2 and imut are time-varying, and the constants a1, a2, m0, m1, and m2 can be moved to the left side of the expression. We then derive the new expression of this criterion by a new variable S:

Equation (19)

For m1 ∈ [5, 15] MJup, a1 = 2.14 au, a2 = 18.62 au, thus S should meet the requirement S < 2.807 to maintain the stability of the system. The calculated maximum S for the rolling-over orbits in Figures 12(a), (c), and (f) is 2.13, which is definitely within the stability criterion. In Figures 12(b), (d), and (f), the calculated maximum S is 2.64, providing stable cases using Equation (19). Hence, the orbital flips in Figure 12 are stable, without the chaotic excitation of the binary's eccentricity and inclination.

Comparing with previous work on the stability of the γ Cep system, Haghighipour (2004) conducted an extensive numerical study of the orbital stability of the γ Cep system and suggested that the system could remain steady for imut ∈ [0°, 60°] and e2 < 0.5. This condition is also supported in this work, since we constrain the eccentricity e2 ∈ [0.35, 0.45], which is well consistent with our simulation results. Taking Figure 12(b) as an example, the critical value of the initial imut that is necessary for orbital flips to occur is lower than 60°, indicating a stable initial status, according to Haghighipour (2004), with regular oscillations of e1 and imut causing the system always to return to a stable situation. Satyal et al. (2013) examined the stability and quasiperiodicity of γ Cep, and explored the orbital stability for various inclinations and binary eccentricities e2 through the reliability comparison of chaos indicator. They demonstrated that the planet γ Cep Ab can remain stable for e2 being as high as 0.6 or for i1 ≤ 25°. In this work, for rolling-over orbits, we demonstrate that γ Cep Ab can also remain stable when e2 ∼ 0.4 and imut ∼ 90°. We further confirm in this work that there is a great possibility of the flipping cases of γ Cep Ab being locked into a Kozai resonance, based on the surfaces of section in Figure 12 and the long-term stability criterion S.

From the simulation results, we also see that the planetary eccentricity could be stirred up to 0.9999, due to secular perturbation from the binary, as close approaches may eject the planet out of the system, so that it would not be observed or its orbit could be shrunk, owing to tides over the evolution. While this process takes too much time to be observed on the timescale of the EKL mechanisms, we employ imut to explore the stability of the system, as the evolutions of e1 and imut are coupled under this scenario. The extreme oscillations of inclination and eccentricity would enhance the rate at which the system would be brought into an unstable situation. Li et al. (2014a) showed that there could be chaotic behavior when the mutual inclination between the inner and outer orbit remained high.

The instability of S-type planets may be induced by large eccentricity excitations. The planet may fall into the region of Roche lobe and merge into the primary, or move away from the inner binary system (Eggleton et al. 1998; Kiseleva et al. 1998; Ford et al. 2000).

5. Maximum Mutual Inclinations of General S-type Systems

After exploring crucial issues about the inclination excitation of γ Cep Ab, we attempt to reveal the dynamical features of general S-type planets under the octupole-level secular resonance. For the secular evolution of general S-type planets, dynamical evolution should be extensively studied in a wider parameter space, thus more parameters are accounted for in this section.

As the distribution of the SMAs of S-type planets has been introduced in Section 1, m1 refers to the mass of the S-type planet, while m2 denotes that of the stellar companion. For general S-type planets in close binary systems, the parameter space of m1, a1/a2, e1, and e2 should be enlarged. In this section, the mass of the primary star is set to be 1 M and the mass of the secondary companion is assumed to be 0.3 M, on the basis of the average stellar masses of detected star binary–hosting S-type planets (Raghavan et al. 2010; Moe & Di Stefano 2017).

In this section, we do not discuss the effect of the initial inclinations; rather, we mainly focus on whether the S-type planets in these systems may have experienced orbital flips under the relative conditions derived from the previous sections of this work, therefore we take the initial mutual inclination imut as 50°, which is a bit larger than the critical inclination of 39° for orbital flips in the classical KL theory. Moreover, we let e1 and e2 uniformly distribute over 0.0 ∼ 0.8, and yield a grid of 16 × 16. For each set of initial parameters, we integrate the system for 100 Myr. The argument of periastron is set as g1 = 5° and g2 = 0°, with longitudes of node h1 = 180° and h2 = 0°.

In order to show the evolution of γ Cep Ab from the current observed position, and to enlarge the parameter spaces of e1, e2 to present more general results, Figure 13 plots the distribution map of the maximum imut for m1 = 11 and 15 MJup for a1 = 2 au, a2 = 20 au. Both panels show that the orbital flips occur when e1 and e2 are both larger than 0.2, or e1 < 0.2 and e2 > 0.3. We mark the orbital position of γ Cep Ab with the yellow pentagram in Figure 13(b). The flip constrains of e1 and e2 reveal that γ Cep Ab still has a great possibility of maintaining the flipping orbit. We use the new variable pflip to describe the flip possibility of a total of 256 runs in each panel: in Figure 13(a), pflip = 0.301, while in Figure 13(b), pflip is calculated to be 0.305, and the difference is mainly induced by the tiny changes in the flip cases in the region of e1 < 0.3.

Figure 13.

Figure 13. Distribution of the maximum imut for S-type planets, with a1 = 2 au and a2 = 20 au. Panel (a): m1 = 11 MJup. Panel (b): m1 = 15 MJup. The color bar on the right represents the maximum value of imut over 100 Myr.

Standard image High-resolution image

As we mentioned in Section 3.2, the SMA ratio a1/a2 is the specific element for evaluating the strength of the octupole-level effect when m0, m1, and e2 are fixed. We further set a1/a2 as a new independent variable, we choose a1 = {1, 2, 5, 10} au and a2 = {10, 20, 50, 100} au, which are selected from the "Target Area" in Figure 1, thus a1/a2 = {0.01, 0.02, 0.04, 0.05, 0.1}. Additionally, we assume the S-type planetary mass to be 1 MJup, according to the detected average of the minimum masses of S-type planets.

Figure 14 shows the distribution of the maximum orbital inclination in the parameter space for general S-type planets with masses of 1 MJup. We show the value of the flip possibility pflip in the lower left corner of each panel. For panels (d), (g), (i), and (j) in Figure 14, with a1/a2 = 0.1, the critical eccentricities for the orbital inverting are e2 ∼ 0.3 when e1 ∼ 0, and e2 ∼ 0.2 when e1 > 0.4. For panels (c) and (h) in Figure 14, with a1/a2 = 0.05, the critical eccentricities for the orbital turning over are e2 ∼ 0.5 when e1 ∼ 0, and e2 ∼ 0.4 when e1 > 0.5; the pflip in these two panels are both larger than 0.1. It should be noticed that in Figure 14(f), with a1/a2 = 0.04, the critical e1 and e2 for the flips are both 0.1 higher than those in Figures 14(c) and (h), with pflip < 0.1. For the remaining plots with a1/a2 = 0.01 or 0.02, the flip regions in the parameter space gradually disappear.

Figure 14.

Figure 14. Distributions of the maximum imut for S-type planetary systems with various initial SMA a1 and eccentricities e1, e2. Each diagram has specific initial parameters a1 and a2. The color bar on the right represents the maximum value of imut over 100 Myr, while the regions in dark red represent orbits with imut excited up to 180°.

Standard image High-resolution image

We first conclude from Figure 14 that the orbital flips occur more easily with the increase of a1/a2, which scales the octupole strength, with the flip possibility rising simultaneously. The other major point to note in Figure 14 is that when a1/a2 is fixed, the regions of the orbital flips in the panels on the diagonal get larger with the decrease of a2, since the gravitational perturbation from the secondary star is also getting stronger.

Comparing the distribution maps of the maximum imut for m1 = 11 and 15 MJup in Figure 13 and that for m1 = 1 MJup in Figure 14(i), we notice that the flip possibilities in these three plots are 0.301, 0.305, and 0.297, respectively, thus the number of flipping cases changes only a little with planetary mass. Anderson et al. (2016) also investigated the ratio of all likely consequences of the inward migration of giants in stellar binaries under the octupole perturbation and the tidal dissipation. They found that the fraction of systems giving rise to either hot Jupiter formation or tidal disruption is constantly 11%–14%, having little variation with planetary mass, stellar type, or tidal dissipation strength. Nevertheless, we still obtain some new findings relating to the variation of planetary masses. As the planetary mass increases, the original flips in the region of e1 ≤ 0.1 and e2 ∼ 0.6 disappear, while there new flips emerge in the region of e1 ∈ [0.2, 0.3] and e2 ∼ 0.8.

In Michtchenko & Malhotra (2004), the theoretical analysis of the secular dynamics for three-body systems was performed within the space (e1, e2), and the phase space structure depended upon the ratios of the planetary masses and their SMAs. After presenting analysis for a wide range of planetary masses and SMA ratios, they showed that when both the mass and SMA ratios are far from unity, the domains of the oscillation orbits decrease. In comparison, our results in Figures 13 and 14 are consistent with those of Michtchenko & Malhotra (2004). In Figure 12, we find that the reverting orbits are almost identical to the oscillating cases, thus the magnitude of the flip possibility pflip (Figures 13 and 14) could be expressed as the region within the (e1, e2) plane that is dominated by oscillations. Thus, we conclude that the flip possibility pflip goes down with decreasing a1/a2 and m1/m2.

We further calculate the orbital flip maps of some specific potentially inclined S-type planets, employing their real minimum planetary masses. Here, we mark out possible orbital positions of the S-type planets HD 19994 Ab (Mayor et al. 2004) and HD 196885 Ab (Chauvin et al. 2011) in Figure 14, since the emergence of the flipping cases changes little with planetary mass for e1 ≥ 0.3. The yellow diamond in panel (a) approximately gives the observed parameters for the S-type planet HD 19994 Ab. HD 19994 Ab was discovered in 2003 by RV measurements, with a minimum mass ${m}_{1}\sin {I}_{1}=1.68$ MJup, a1 = 1.42 ± 0.01 au, and e1 = 0.3 ± 0.04. The binary HD 19994 AB consists of a primary of m0 = 1.34 M and a secondary of m2 = 0.35 M, a2 = 100 au, and e2 = 0.26. Figure 14(a) indicates that HD 19994 Ab has a high probability of retaining mutual inclination below 60° for the minimum mass.

Aside from HD 19994 Ab, we also discuss the flip possibility of HD 196885 Ab, with ${m}_{1}\sin {I}_{1}=2.98$ MJup, a1 = 2.0 au, and e1 = 0.48. HD 196885 AB binary consists of a primary star with m0 = 1.33 M and a secondary with m2 = 0.55 M and a2 = 23 au. The problem is that we do not know the eccentricity of the secondary star. Here, we make some assumptions about e2: if e2 ≥ 0.65, then the orbit of HD 196885 Ab will roll over for m1 = 1 MJup, otherwise, it will not flip. Since we have found that the planetary mass does not change the orbital flip possibility for e1 ≥ 0.3, this assumption remains valid as m1 increases.

6. Conclusions and Discussion

In this work, we employ the nonrestricted EKL mechanism to shed light on the secular evolution of the inclined S-type planet γ Cep Ab. Using a wide range of parameters of SMA, eccentricity, and planetary mass, we perform numerical simulations in relation to the octupole-level effects to extensively investigate the orbital flip possibility of potentially inclined S-type planets in general systems. Here, we summarize the major results as follows:

  • 1.  
    We first derive the posterior distributions of the orbital parameters of γ Cep Ab and the star companion in the γ Cep system with more accurate estimation uncertainties in the N-body model, using the MCMC ensemble sampler and the initial parameters independently. The minimum planetary mass is further estimated.
  • 2.  
    We then employ the EKL mechanism to explain the origin of the high inclination of γ Cep Ab. With initial conditions selected from the (e1, imut) plane, we show that when m1 = 15 MJup, it is easier for γ Cep Ab to reach the target imut over 113° when the initial imut < 60° and e1 < 0.7. Our investigation further indicates that relatively small values of ${ \mathcal H }$ and low initial imut may lead to a slightly longer timescale for the first orbital flip. The libration and circulation regions in the (e1, g1) plane and Poincaré surfaces of section, as well as the secular stability criterion, confirm that there is a great possibility for the flipping orbits of γ Cep Ab to remain stable.
  • 3.  
    This work further extends the application of the EKL mechanism to general S-type planets. We take m1, a1, and e2 as independent variables, and the orbital flips can be observed in the selected space of a1 and a2. The most intense orbital inclination excitation occurs when a1/a2 = 0.1 and e2 ∼ 0.8. As the planetary mass increases, the original flips at e1 ≤ 0.1 and e2 ∼ 0.6 are suppressed, whereas the new flips emerge for e1 ∈ [0.2, 0.3] and e2 ∼ 0.8, with little effect on the total flip possibility.

In this study, we mainly focus on the extremely high mutual inclination or transformation between prograde and retrograde orbits in binary systems over the timescale of secular resonance. Note that the EKL mechanism can trigger high eccentricities of the planet, and the final orbit of the planet is closely related to e1 and tidal dissipation. The star companion can drive the planet to approach the central star through eccentricity excitation in the secular chaos, and the tidal dissipation will eventually circularize them into hot Jupiters (Wu & Murray 2003; Wu & Lithwick 2011). However, there still remain several open questions for forthcoming investigation, e.g., with what kinds of mechanisms are the excited eccentricity of the planet reduced to approximately zero.

Actually, the comprehensive effects of the EKL mechanism and tidal theory will be explored in a future study. We will construct a double-averaged model that involves the octupole-level secular resonance and the equilibrium tides of spinning planets. The extreme evolution of the periastron distance induced by secular resonance could significantly vary the tidal dissipation timescale. Our simulations for hot Jupiters in binary systems indicate that the eccentricity of the planet can decrease from 0.8 to 0 over less than 104 years. Recent studies have also shown that the combination of high-eccentricity migration and the KL scenario can be applicable to the formation of diverse exoplanets (Petrovich et al. 2019; O'Connor et al. 2021).

Anderson et al. (2016) have explored inward migration for giant planets in stellar binaries via EKL and showed that the gas giants can be circularized as hot Jupiters. For S-type planets falling into the region dominated by the tide, secular evolution in combination with tidal dissipation will provide a more comprehensive picture of dynamical evolution. Alternative studies suggest that the mean motion resonance as mixed with secular resonance as well as other scenarios can play a vital role in the evolution of giant planets in binary systems or those that are hosted by single stars (Haghighipour 2006; Bazsó et al. 2017; Liu & Ji 2020). Hence, extensive investigations with additional high-precision observations from space-based/ground-based telescopes will be necessary to better understand the complicated dynamics and formation of S-type planets.

We thank the anonymous referee for the constructive comments and suggestions that significantly improved the original manuscript. This work is financially supported by the National Natural Science Foundation of China (grant Nos. 12033010 and 11773081), the B-type Strategic Priority Program of the Chinese Academy of Sciences (grant No. XDB41000000), and the Foundation of Minor Planets of the Purple Mountain Observatory.

Footnotes

Please wait… references are loading.
10.3847/1538-3881/ac8f4c